Next Article in Journal
Identification of Novel miRNAs and Their Target Genes in the Response to Abscisic Acid in Arabidopsis
Previous Article in Journal
Regenerative Effects of Hypoxia Primed Flowable Placental Formulation in Muscle and Dermal Injury
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Basic Helix-Loop-Helix (bHLH) Transcription Factors Regulate a Wide Range of Functions in Arabidopsis

Chinese Education Ministry’s Key Laboratory of Western Resources and Modern Biotechnology, Key Laboratory of Biotechnology Shaanxi Province, College of Life Sciences, Northwest University, China P.R. 229 North Taibai Road, Xi’an, Shaanxi 710069, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(13), 7152; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22137152
Submission received: 25 May 2021 / Revised: 29 June 2021 / Accepted: 29 June 2021 / Published: 1 July 2021
(This article belongs to the Section Molecular Plant Sciences)

Abstract

:
The basic helix-loop-helix (bHLH) transcription factor family is one of the largest transcription factor gene families in Arabidopsis thaliana, and contains a bHLH motif that is highly conserved throughout eukaryotic organisms. Members of this family have two conserved motifs, a basic DNA binding region and a helix-loop-helix (HLH) region. These proteins containing bHLH domain usually act as homo- or heterodimers to regulate the expression of their target genes, which are involved in many physiological processes and have a broad range of functions in biosynthesis, metabolism and transduction of plant hormones. Although there are a number of articles on different aspects to provide detailed information on this family in plants, an overall summary is not available. In this review, we summarize various aspects of related studies that provide an overview of insights into the pleiotropic regulatory roles of these transcription factors in plant growth and development, stress response, biochemical functions and the web of signaling networks. We then provide an overview of the functional profile of the bHLH family and the regulatory mechanisms of other proteins.

1. Introduction

Plants frequently encounter a wide array of environmental stresses, including drought, cold, salinity, UV radiation, ozone and other abiotic stresses through their lifecycle, which severely affect plant growth, development and crop yield. To cope with those problems, plants adapt to the different harsh conditions through transcription factors that regulate the expression of a series of target genes. The basic helix-loop-helix (bHLH) protein family is one of the largest transcription factor gene families in Arabidopsis thaliana, and contains a highly conserved amino acid motif that can be found throughout eukaryotic organisms. They were originally identified in animals, and found to play important roles in diverse regulatory networks, via the functions of DNA binding and dimerization [1,2]. The binding and dimerization networks of those proteins also have diverse functions in plants, including in multiple aspects of plant growth and development, which were characterized in recent years.
The bHLH superfamily, named after its highly conserved alkaline/helix-loop-helix domain, is the second largest transcription factor family and is widely present in animals and plants [3,4]. The bHLH domain is composed of approximately 60 conserved amino acid residues and consists of two conserved motifs, namely, a basic region and helix-loop-helix region (HLH region). The basic region participates in DNA binding to E-box (usually CANNTG) or G-box (CACGTG) motif in their target genes, and the HLH region, which is composed of two alpha helices containing hydrophobic residues, is required for dimerization to change the expression of target genes involved in various signaling pathways. Based on their phylogenetic relationship and DNA binding function, the bHLH members in eukaryotes have been classified into six main groups (named A, B, C, D, E and F) [5]. In brief, Group A can specifically bind to the E-box core sequence, Group B prefers to bind the G-box, and Group C binds to the ACTTG or GCGTG sequences [6,7]. Group D lacks a typical basic region and mainly participates in heterodimerization with other bHLH family proteins [8]. Group E can preferentially bind to the CACGNG sequence, and the Group F members can bind to specific DNA target sequences [9,10,11]. At least 162 members of bHLH proteins have been identified in Arabidopsis [12,13,14], which could be clustered into 12 distinctive subfamilies, namely bHLH I to XII [13,14]. Though bHLH proteins in Arabidopsis are quite divergent, the most abundant type of Arabidopsis bHLHs are more closely related to the phylogenetic group B of eukaryotic bHLHs, which is proposed to be the ancestral bHLH type [14,15]. Arabidopsis also contains some atypical bHLH proteins, which lack the basic region and are referred as AtHLHs. However, they do not share a close sequence similarity to the group D proteins in animals [4,14].
As a superfamily, bHLHs have been characterized in Arabidopsis over the years, and members of this family members play vital roles in plant growth, development, light signal transduction and stress responses [16,17,18,19,20]. They are also involved in the crosstalk of hormones signaling, including abscisic acid (ABA), jasmonic acid (JA), brassinosteroid (BR), salicylic acid (SA) and ethylene (ET) [1,4,13,21], and they are pivotal for plant growth and survival in the environment.
To date, many excellent reviews have covered the roles of bHLH transcription factors in plants in different contexts, but an overall summary is not available. Thus, we focused mainly on summarizing the knowledge on bHLH functions in various essential processes, such as how bHLH proteins mediate plant growth and development, including flowering time, seed germination and cell fate determination; how they are involved in maintaining iron homeostasis in plants, how they respond to light, stress, phytohormones and the crosstalk among these factors; and how they perform the functions in other processes, such as senescence and anthocyanin biosynthesis (Figure 1 and Table 1).

2. Functions of bHLH Factors in Plant Growth and Development in Arabidopsis

2.1. Roles in Seed Germination

Seed dormancy and germination are critical processes in the lifespan of plants, which are mediated by various external environmental factors such as temperature, light and humidity, and internal factors, such as phytohormones. Two type of phytohormones, ABA and gibberellin (GA), are crucial regulators of seed dormancy and germination. ABA enhances seed dormancy, whereas GA breaks dormancy and promotes germination. The transcription factor bHLH57 is involved in ABA-regulated the seed dormancy process. NCED (9-cis-epoxycarotenoid dioxygenase) is considered to catalyze the first dedicated step in ABA biosynthesis [22], ODR1 (reversal of rdo5, an ortholog of the rice seed dormancy 4 (Sdr4)) protein interact with bHLH57 in a complex to inhibit NCED gene transcription [23,24].
In addition to phytohormones, photoreceptors also regulate seed dormancy and germination. Phytochrome interacting factors (PIFs), a group of typical of bHLH proteins, could interact with diverse groups of factors to integrate external environmental and internal signals, and further control seed germination, shade avoidance and crosstalk of plant hormones and the clock-derived signaling pathway [25,26]. The phytochrome-interacting bHLH protein PIL5 (PIF3-like 5 PIF1/bHLH15) is a repressor of seed germination, acting by reducing GA level in the dark; phytochromes accelerate the PIL5 degradation and increase the level of bioactive GA in seeds upon exposure to light. In addition, PIL5 activates the expression of the ABA synthesis gene and maintains a high level of ABA in the dark to inhibit seed germination [27]. Another light stable bHLH transcription factor SPATULA (SPT) also mediates seed germination, by interacting with PIL5 in response to light and temperature [28,29].

2.2. Functions in the Flowering Time Control

It is critical for plants to manage their flowering time, which is regulated by an intricate network of molecular signaling and controlled by various environmental factors such as photoperiod and temperature. The CONSTANS (CO) protein functions as an essential component for transforming biological clock signals into flowering signals to initiate plant flowering. Four bHLH related proteins (flowing bHLHs), namely, FBH1, FBH2, FBH3 and FBH4, are transcriptional activators of the CO gene; they bind to the E-box cis-elements in the promoter of CO and then positively regulate the CO-mediated flowering time [30]. Cryptochromes (CRY1/2) are blue light receptors that inhibit hypocotyl elongation and control floral initiation [31]. CRYPTOCHROME-INTERACTING BASIC-HELIX-LOOP-HELIX 1 (CIB1) is a CRY2-interacting bHLH protein that promotes the FLOWERING LOCUS T (FT) transcription; CRY2, CIB1 and CO can form a protein complex in response to blue light and then promote floral initiation [32,33].
Recently, a bHLH transcription factor, named as NO FLOWERING IN SHORT DAY (NFL) was shown to be necessary for the promotion of flowering specifically under short-day (SD) conditions, and nfl mutants did not flower under SD conditions but were similar to wild type under long-day (LD) conditions [34]. Additionally, three myelocytomatosis (MYC) proteins redundantly regulate flowering under both LD and SD conditions. Although MYC4 (bHLH4) is more important in mediating flowering than MYC3 (bHLH5), MYC2/3/4 are required in the JA pathway for regulating the flowering time by inhibiting FT transcription in Arabidopsis [35]. Moreover, the Brassinosteroid Enhanced Expression1 (BEE1, bHLH44) protein is stabilized under blue light; this protein is an integrator regulating photoperiodic flowering [36].
In Arabidopsis, temperature also affects the flowering time, and high temperature (29 °C) not only induces rapid hypocotyl elongation but also results in early flowering. In addition, the central integrator bHLH transcription factor phytochrome-interacting factor 4 (PIF4) can accelerate the flowering process by activating FT at high temperature conditions [37]. In addition to regulating the flowering time, bHLH family members are involved in the flower organ development. SPATULA (SPT) encodes a bHLH transcription factor and plays a role in floral morphogenesis processes as previously described [38].
In summary, the flowering time is strictly controlled by an intricate network, and bHLH family members act together with many other proteins to allow plants to flower at suitable time and under favorable environments.

2.3. Functions on Cell Fate Determination

Plants process a cell determination mechanism for formatting specific cell types, and this system relies on the expression of different genes in a proper spatiotemporal manner. Many transcription factors participate in this critical process, including a number of bHLH transcription factors, which play essential roles in the root and shoot cell fate determination [39]. During the root development, two types of cells arise from the epidermis: root hair cells and non-hair cells. Root hair cells produce an outside long tubule from a specialized root surface area and have functions in absorbing nutrition and water, and also in interacting with microbes.
The bHLH VIIIc subfamily member, bHLH83 (root hair defective 6, RHD6), is a transcription factor that has fewer root hairs and functions in the inhibition root hair formation in cortical cells [13,39], and another bHLH protein, ROOT HAIR DEFECTIVE 6-Like 1 (RSL1), functions redundantly with RHD6 to modulate the root hair development [40]. RSL2, a homolog of RSL1, is also required for the normal root hair formation, which is controlled by RHD6 and RSL1 [41]. Additional genes, such as RSL3/4 and LRL3 (Lj-RHL1-LIKE3) also act as downstream of RHD6 and RSL1 to promote the root hair and cell differentiation [41,42,43]. Moreover, RSL4 is suppressed by GLABRA2 (GL2) through ZINC FINGER PROTEIN1 (ZP1) [44]. GL2, a member of homeodomain-leucine zipper (HD-ZIP) protein, affects the epidermal cell fate including in trichomes, root hair and the seed coat. It represses the transcription of RHD6 to inhibit hair formation in N-cells (non-root hair cells), leading to the expression of non-hair genes [42,45]. The bHLH transcription factor GL3 (GLABRA3, bHLH1) has similar functions, regulating its own transcription and downstream target genes to trigger the trichome initiation pathways, ENHANCER OF GLABRA3 (EGL3, bHLH2) plays partially redundant roles with GL3 in root hair [46,118].
Furthermore, ET is involved in root hair initiation and elongation, and ETHYLENE INSENSITIVE 3 (EIN3) and its homolog EIN3-like1 (EIL1) can coordinate with RHD6/RSL1 to upregulate RSL4 and then participate in root epidermis development [119]. Another bHLH complex named MYB-bHLH-WD40 was proposed to regulate the guard cell and root hair differentiation [47].
In addition to their roles in the root epidermis formation, the bHLH family is also involved in other cell fate related processes, including stamen and stomatal development. The IIIe bHLH transcription factor MYC5 (bHLH28) has redundant functions with MYC2 (bHLH6), MYC3 (bHLH5) and MYC4 (bHLH4), which interact with MYB21 and MYB24 to form the bHLH-MYB complex to regulate the stamen development [48]. In addition, FAMA (bHLH97) is specifically expressed in the stomata and has functions in halting proliferative division and promoting guard cells [120].
Overall, the establishment of this specification mechanism is complicated and also influenced by several plant growth regulators, including hormones such auxin (indole-3-acetic acid, IAA), ET, JA [55,119].

3. Functions in Environmental Response

3.1. Functions in Plant Mineral Nutrition and Abiotic Stress

Iron is an indispensable mineral micronutrient for photosynthesis and respiration in plants. Low concentration of effective iron in soil leads to the iron deficiency in plants, conversely, too much iron can have a toxic effect. To cope with this problem, plants have developed a series of sophisticated regulatory systems to control the iron uptake and maintain the Fe homeostasis. Among the regulatory mechanisms, bHLH transcription factors play important roles in this complicated process [49,92].
Four closely related clade Ib bHLH genes, namely, bHLH38, bHLH39 [50], bHLH100 and bHLH101 [121], have been characterized as key transcription factors regulating the Fe-deficiency response in Arabidopsis, and their transcript levels were shown to be up-regulated under Fe deficiency [50,121]. A single loss-of-function mutant of these genes did not show an obvious change in phenotype compared with wild-type because of functional redundancy in this subfamily. A triple loss-off function mutant of bHLH39, bHLH100 and bHLH101 showed severe leaf chlorosis only under Fe-deficient conditions [51]. bHLH29 (FER-like iron deficiency-induced transcription factor, FIT), an ortholog of FER (from the T3238FER line), acts as a central regulator to iron deficiency inducible genes in iron uptake and homeostasis [52,53,54]. Research has indicated that FIT interacts with bHLH38, bHLH39 [50], bHLH100 and bHLH101 [121] to form heterodimers that activate the transcription of genes involved in the Fe uptake system. Recent research has shown that bHLH121 can form homo or heterodimers and directly bind the FIT promoter to regulate its expression [122]. In addition, cadmium (Cd) stress is a severe heavy metal stress that limits plant growth, and FIT/bHLH38 or FIT/bHLH39 heterodimers can respond to Cd tolerance by affecting the transcriptional expression of downstream genes [56].
Four IVc bHLH transcription factors (bHLH34, bHLH104, bHLH105/ILR3 - IAA-LEUCINE RESISTANT3, and bHLH115) form homo- and heterodimers to regulate the Fe deficiency response and the Fe homeostasis [57,58,59,92]. Knockout of bHLH104 reduced Arabidopsis tolerance to Fe deficiency and suppressed the activation of Fe deficiency-inducible genes, whereas overexpression had the opposite effect. The bHLH034, homolog of bHLH104, also positively regulated the Fe deficiency response. The bHLH105, also named IAA-LEUCINE RESISTANT3 (ILR3), interacted with bHLH104, which plays an important role in maintaining the Fe homeostasis. Notably, a triple mutant deficient of bHLH34/104/105 is sensitive to Fe deficiency [59].
In the case of FIT, its function is required to regulate Fe uptake responses; bHLH47 (POPEYE, PYE) is another bHLH transcription factor that plays a major role in the Fe homeostasis. PYE interacts with bHLH104, bHLH105 and bHLH115 to bind to the promoters of several Fe response genes and repress their activity [60,61,62,123]. Interestingly, the bHLH104, bHLH105 and bHLH115 protein levels are posttranslationally controlled by BRUTUS (BTS), a hemerythrin E3 ligase, via the proteosomal degradation to negatively regulate the bHLH protein stability [61,124]. It was reported that the BTS paralogs, BTS LIKE1 (BTSL1) and BTSLIKE2 (BTSL2) act redundantly as negative regulators of the Fe starvation response, and the transcript levels are also controlled by FIT [125].
In most plants, Fe(II) is taken up by IRON REGULATED TRANSPORTER 1 (IRT1) in the root plasma membrane from soil under iron deficiency [63,126,127]. Moreover, ferric reduction oxidase 2 (FRO2) transfers electrons across the plasma membrane for iron reduction at the root surface. These are key genes regulating ferric reduction and iron transport in plants, and their transcription levels are upregulated by FIT and bHLH Ib heterodimers [50,121]. Furthermore, bHLH11 acts as a negative regulator to modulate Fe levels in Arabidopsis, and bHLH11 is highly expressed in roots and its expression is upregulated after plants are transferred to Fe sufficient conditions [64].
In addition to regulating the iron response, it was reported that bHLH104 loss-of-function mutants were sensitive to Cd stress, and the Cd tolerance was enhanced upon overexpression of bHLH104 [20]. In other words, bHLH104 positively regulates both Cd tolerance and Fe deficiency tolerance in Arabidopsis.

3.2. Functions in Stress Responses

Inducer of CBF Expression (ICE1, bHLH116) encodes an MYC-type bHLH transcription factors that directly binds the promoter of the C-Repeat Binding Factor/Dehydration-Responsive-Element-Binding protein (CBF/DREB1) protein-encoding gene and activates its transcription to improve cold tolerance through an ABA independent pathway [64]. The regulation is controlled by the binding of various downstream target cold and dehydration response genes that contain specific binding sites [65,128,129,130].
Other transcription factors, such as AIF2 (also named RITF1) and its partner RSA1 (SHORT ROOT IN SALT MEDIUM 1) regulate several gene for detoxification of reactive oxygen species (ROS), which are triggered by the salt stress [66].
Some proteins, such as bHLH122, can directly bind to the CYP707A3 promoter to repress its expression and enhance the ABA content, but these proteins were not induced by ABA treatment, indicating that they may function in ABA-independent or ABA-dependent ways [22,67].
As direct target genes of PIF4, JUNGBRUNNEEN1 (JUB1) and ORESARA1 (ORE1) have vital roles in salt stress response. PIF4 directly regulates ORE1 and JUB1, and is involved in regulating salt stress response [68,69,70].

4. Functions in the Response to Light and Phytohormones

Under specific conditions, multiple phytohormones and environmental factors are in constant crosstalk with each other to affect the plant growth and development. Based on previous studies, GA, IAA, BR, ET and light are usually considered to promote cell expansion in plant growth, while ABA, JA and SA are normally involved in the response to biotic and abiotic stresses. bHLH proteins that respond to plant hormones and environmental factors participate in various processes. They are able to interact with each other to cooperatively or antagonistically modulate the plant growth.

4.1. Response to Light Signaling

Light is an important factor that influences the plant growth and development, and many bHLH transcription factors are reported to participate in this process by inducing the expression of related downstream genes and various light-mediated effects.
PIFs interact with the active form (Pfr) of phytochrome to modulate growth, including in response to environmental signals such as light and stress and via other signaling pathways [73]. The PIF family belongs to the bHLH superfamily VII of transcription factors, which play central roles in the regulation of light signaling. To date, 15 PIF members have already been identified, and 7 PIF members have been shown to bind the Pfr form of phyB in Arabidopsis, while other members do not interact with the light-activated phytochrome [13,14,131,132,133]. Each PIF has individualized or redundant biological functions with other PIF proteins during various responses.
In the dark, four PIFs (PIF1/PIL5, PIF3, PIF4 and PIF5/PIL6) directly interact with Pfr to promote skotomorphogenesis by repressing photomorphogenesis. Loss-of-function of pif mutant showed phenotypes of reduced hypocotyl elongation. PIFs also interact with diverse groups of transcription factors to integrate external environmental and internal signals, including in seed germination, shade avoidance and crosstalk among plant hormones and clock-related signaling pathways [24,25]. A number of genes have been confirmed to be direct targets of PIF, thus mediating downstream light signaling networks through the PIFs.
PIF3 (bHLH8), the first identified member in the PIF family, acts as a regulator in the seedling de-etiolation and modulates both positive and negative response to phytochrome-mediated signaling [24,71]. PIF3 can also heterodimerize with the atypical basic bHLH protein HFR1 (long hypocotyl in far-red, bHLH26) to modulate phyA signaling [75]. In addition, PIF3 is involved in plant freezing response as a negatively regulator which can interact with EBF1 (EIN1-BINDING F-BOX1) through the CBF (C-REPEAT BINDING FACTOR) pathway [72].
PIF4 is a key transcription factor in the light signaling pathway; it interacts selectively with Pfr and negatively regulates phyB signaling in Arabidopsis [74], Phytochrome Rapidly Regulated 1 (PAR1, bHLH165) and its homolog PAR2 (bHLH166) lack the ability to directly interact with phytochromes that are rapidly induced by shade [76,77], while PAR1 directly interacts with PIF4 to form a heterodimer to inhibit PIF4 function in cell elongation [78]. Moreover, growing evidence indicates that PIF4 acts as central regulator, coordinating plant response to multiple environmental signals [134].
The PIL (PIF3-like proteins, PILs) family has six members, designated as PIL1, PIL2, PIL3, PIL4, PIL5 and PIL6. PIF1/PIL5 (bHLH15) is a component that negatively regulate the chlorophyll biosynthetic pathway, seeds germination and inhibit hypocotyl elongation in the dark. The activity of PIF1 was repressed by phyA and phyB in light, and regulated ABA signaling [79,80,81,82]. In addition, other PIFs, such as PIF5 (PIL6), PIF6 (PIL2) and PIF7, have been shown to interact with the Pfr form of phyB or phyA, which are involved in phytochrome signaling [83,84,85].
Taken together, the bHLH proteins PILs/PIFs are proposed to form heterodimers to regulate bHLH network activity and are central components that integrate multiple signals in response to light.

4.2. Functions in JA Signaling Pathway

The phytohormone jasmonate acid (JA) plays a vital role in the plant development and the response to various stresses. The presence of JA triggers the key protein jasmonate ZIM-domain (JAZ) to interact with Coronatine Insensitive 1 (COI1), part of the SCFCOI1 ubiquitin E3 ligase complex. Then, JAZ proteins were degraded by the 26S protease, resulting in multiple transcription factors free from JAZ-mediated repression and further activating downstream JA-mediated responsive genes. The bHLH transcription factor MYC2 forms a signaling module COI1/JAZs/MYC2 to participate in the JA-induced signaling pathway. The homologs of MYC2, MYC3 (bHLH5) and MYC4 (bHLH4) are known to form homodimers/heterodimers, and they can also bind to the G-box involved in the JA signaling pathway but exhibit gene redundancy with MYC2 [86,87].
The bHLH transcription factor GLABRA3 (GL3, bHLH1) can form a WD-repeat/bHLH/MYB complex with TRANSPARENT TESTA GLABRA1 (TTG1) and the R2R3-MYB transcription factor GLABRA1 (GL1), which is repressed by JAZs and DELLAs, is responsible for trichome initiation [88,89,90,91].
Four subgroup IVa bHLH transcription factors (bHLH18, bHLH19, bHLH20 and bHLH25) can be induced by JA and inhibit the transcription of the FIT and Ib bHLH genes, which have been suggested to function redundantly in JA-mediated FIT protein degradation in the presence of JA or under iron deficient conditions via the 26S proteasome pathway [92].
The jasmonate-activated transcription factor MYC2 has also been found to interact with the key component in the ET signaling pathway EIN3 and its close homolog EIL1 to repressed its DNA binding activity and affect hook formation [135], suggesting that jasmonate and ET have antagonistic functions during apical hook development.

4.3. Functions in IAA Signaling Pathway

The plant hormone IAA has multiple roles in the plant growth and development, such as in cell division, cell elongation and cell differentiation, which are affected by the regulation of IAA response genes [93]. BIGPETALp (BPEp, bHLH31) is a bHLH transcription factor that can interact with AUXIN RESPONSE FACTOR8 (ARF8) to influence cell expansion and petal growth [94], PRE6 is a target of ARF5 and ARF8 that negatively regulates auxin response genes in Arabidopsis [95].

4.4. Roles in ABA Signaling Pathway

Plant are constantly under extrinsic abiotic/biotic environmental stresses, including cold, drought, high salinity, pathogen and extreme temperature. In response to these diverse stresses, plants have evolved sophisticated adaptation mechanisms. To date, several bHLH transcription factors have been reported to mediate abiotic and biotic stress signaling pathways to regulate plant responses in Arabidopsis in different ways.
The plant hormone ABA plays a central role in a variety of physiological processes and environmental response involved in plant growth, including responses to drought, cold, heat and salinity stresses [136,137,138]. Several bHLH transcription factors, such as bHLH112, MYC2, AIB, AtAIG1, bHLH129 and bHLH92, have been reported to be involved in the regulation of ABA signaling via these processes [96,98,99,100].
A loss-of-function mutant of bHLH112, a transcriptional activator, displayed a late-flowering phenotype under long day conditions in Arabidopsis, and its transcript level was correlated with salt and drought tolerance. bHLH112 regulates gene expression by binding to E-box and GCG-box motifs in the gene promoters and then mediate multiple physiological response to improve stress tolerance [96,97].
ABA-inducible bHLH-type transcription factor (AIB) and ABA-inducible gene (AIG1) are ABA-induced genes, the proteins contain a bHLH type DNA binding domain and play a positive role in ABA signaling in Arabidopsis [98,99]. MYC2 and MYB2 have been shown to be transcription activators that function in the ABA signal transduction pathway by directly regulating the expression of ABA response genes [17]. In addition, MYC2 can also be involved in JA [139,140,141].
The expression level of bHLH129 is reduced upon exogenously application of ABA, and bHLH129 regulates the expression of several ABA signaling component genes [100] to promote root elongation.
PREs are involved in plant growth and development and are also involved in the regulation of ABA mediated salt responses in Arabidopsis. Some PREs are ABA responsive genes; their expression levels are decreased under ABA treatment, and this response has functions in regulating plant growth and environmental stimuli [101].
ABA induces stomatal closure and changes the expression of numerous genes to adapt to drought stress, ABA-responsive kinase substrate (AKS1), also a bHLH transcription activator, is inhibited by ABA through phosphorylated to form monomer by SNF1-related protein kinase 2 (SnRK2) in Arabidopsis guard cells [142,143].
In addition, the BEE transcription factor family members (BEE1/2/3) regulate plant responses to abiotic stress. BEE genes are strongly repressed by ABA and are redundant negative regulators of physiological responses to abiotic stress, whereas the BEE2 dimerized protein IBH1 is a positive modulator that improves salt and drought tolerance [102]. NaCl-induced expression of bHLH92 confers tolerance to salt and osmotic stress which is partially dependent on ABA and SALT OVERLY SENSITIVE 2 (SOS2) [103].

4.5. The Cross-Talk between Light and Phytohormones

Some bHLH transcription factors are involved in signal transduction networks mediated by plant hormones. For example, several genes are involved in light and GA signaling pathways, such as PIL5 and SPT, which also belong to the bHLH transcription factor family [27,80]. Moreover, many bHLH transcription factors have been reported to be involved in the regulation of BRs, ABA and IAA signaling pathways. Paclobutrazol Resistance (PREs) proteins are atypical bHLH transcription factors that lack the DNA binding domain but can dimerize with bHLH factors to inhibit DNA binding [107,144]. PREs participate in various hormone-, temperature- and light-responsive signaling pathways to regulate plant growth and development in many ways [104,105,107,145]. Briefly, PRE1 (bHLH136), PRE5 (bHLH164) and PRE6 (bHLH163) are direct targets of PIF4 and are involved in the regulation of the light, GA and BR signaling pathways [25,104]. PRE1 regulates cell elongation in Arabidopsis together with ARF6 [105]. PRE3 (bHLH135), also named Activation-Tagged BRI1 Suppressors 1 (ATBS1), can suppress the BR insensitive 1 (bri1) phenotype, and its mutant shows an auxin-related phenotype [16]. PRE4 (bHLH161, BNQ3) mutant shows a light-related phenotype, including pale-green sepals, decreased chlorophyll levels and late flowering [106]. PRE2 (bHLH134) and PRE6 (bHLH163) are ABA response genes that affect plant sensitivity to ABA, indicating that some PREs are involved in ABA and salts responses [95].
Several atypical bHLH proteins, such as Arabidopsis ILI1 binding bHLH1 (IBH1) and ATBS1 Interacting Factors (AIFs), negatively regulate cell elongation in Arabidopsis [107,108]. Activators for cell elongation (ACEs) can promote cell elongation, while IBH1 interacts with ACEs to inhibit their functions in the induction of cell elongation. Another bHLH protein, homolog of BEE2 interacting with IBH1 (HBI1), is involved in BR-mediated growth to promote cell elongation [104]. Functional analysis showed that bHLH proteins have related functions but with different mechanisms to regulate cell elongation. In these processes, bHLHs proteins are mostly dependent on multiple phytohormones including BR, GA, IAA and light signaling [108,146].
BEE1 (bHLH44), BEE2 (bHLH58) and BEE3 (bHLH50) are functionally redundant bHLH transcription factors for which expression is induced by BL treatment, indicating that they function in the early response to BRs and their expression is required by ABA [18]. In brief, the BEE1 protein is stabilized under blue light, which is an integrator to regulate photoperiodic flowering [109]. BEE2 and CIB1 negatively regulate immunity and are functionally redundant with HBI1 in plant immunity [104,110]. Moreover, BEEs positive regulate shade avoidance syndrome with BES1-INTERACTING MYC-LIKE protein (BIMs) [111,112].

5. Functions in Other Aspects of Plant Biology

In Arabidopsis, bHLH members are also involved in other aspects of plant growth and development, such as senescence and the anthocyanin biosynthesis pathway. These aspects are only briefly discussed here.
In plants, senescence is triggered by developmental and environmental factors. The IIId bHLH transcription repressors (bHLH3, bHLH13, bHLH14 and bHLH17) can bind to the promoter of senescence-associated gene 29 (SAG29) to repress MYC2-activated leaf senescence [47]. PIFs can promote leaf senescence during age-triggered and dark-induced processes [113].
bHLH transcription factors have been identified as key regulators of anthocyanin biosynthesis in various plant species [147]. MYC like bHLH proteins can physically interact with MYB transcription factor and WD40-repeat protein genes to form complexes, known as MBW complex, that control the transcript levels of genes by binding to their promoter regions, thereby regulating anthocyanin in biosynthesis [22,114], and these complexes are also crucial for the regulation of flavonoid pathway [115,116,117].

6. Conclusions and Future Prospects

As the second largest superfamily in Arabidopsis, most bHLH transcription factors are characterized by a signature domain, which consists of approximately 60 amino acids, with an N-terminal basic DNA binding domain and a C-terminal protein interaction domain. Although more than 160 genes have already been predicted to belong to the bHLHs family and are classified into 12 subfamilies [14], relatively few have been characterized according to current studies [13,148]. Recent research studies have demonstrated that bHLHs transcription factors are characterized by their roles in a broad range of plant growth and various developmental processes.
Some transcription factors play the more crucial roles in many aspects of plant development, such as MYCs, PREs and PIFs, and these bHLHs are key transcriptional regulators in the phytohormone crosstalk pathway and multiple biosynthesis pathways. As JAZ targets, MYCs are involved in JA-mediated gene expression and the ABA signaling pathway; they also mediate ET biosynthesis and regulate stamen development and flowering time. PREs respond to GA, BR, temperature and light signaling to positively regulate cell elongation. PIFs are also involved in many other signaling pathways, such as in mediating pathways responsive to light, GA, IAA, ABA, ET and abiotic/biotic stress [73,149,150] (Figure 2).
Overall, in this review, we have compiled the current research on bHLH function and provided a relatively complete overview of the bHLH transcription factor family. This review enriches our understanding of this family and provides new insight into the mechanism by which bHLHs regulate various biological processes. However, there remain many aspects that have not been described because of the complicated crosstalk with other transcription factors. Our understanding of function the bHLH proteins have improved tremendously during the past several years, and the research in this area is well characterized. Future studies will further elucidate the mechanism of how bHLH proteins coordinate the multiple internal and external environments to regulate plant growth.

Author Contributions

Conceptualization, Y.H. and A.F.; writing—original draft preparation, Y.H., X.Z., P.R. and Y.Q; writing—review and editing, Y.Q. and A.F.; supervision, A.F. All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by the National Natural Science Foundation of China to Y.H (31300204), Shaanxi Provincial Science and Technology Department to Y.H (2019JQ-326).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable. No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Murre, C.; McCaw, P.S.; Baltimore, D. A new DNA binding and dimerization motif in immunoglobulin enhancer binding, daughterless, MyoD, and myc proteins. Cell 1989, 56, 777–783. [Google Scholar] [CrossRef]
  2. Massari, M.E.; Murre, C. Helix-loop-helix proteins: Regulators of transcription in eucaryotic organisms. Mol. Cell. Biol. 2000, 20, 429–440. [Google Scholar] [CrossRef] [Green Version]
  3. Duek, P.D.; Fankhauser, C. bHLH class transcription factors take centre stage in phytochrome signalling. Trends Plant Sci. 2005, 10, 51–54. [Google Scholar] [CrossRef]
  4. Feller, A.; Machemer, K.; Braun, E.L.; Grotewold, E. Evolutionary and comparative analysis of MYB and bHLH plant transcription factors. Plant J. 2011, 66, 94–116. [Google Scholar] [CrossRef] [PubMed]
  5. Atchley, W.R.; Fitch, W.M. A natural classification of the basic helix-loop-helix class of transcription factors. Proc. Natl. Acad. Sci. USA 1997, 94, 5172–5176. [Google Scholar] [CrossRef] [Green Version]
  6. Henriksson, M.; Lüscher, B. Proteins of the Myc network: Essential regulators of cell growth and differentiation. Adv. Cancer Res. 1996, 68, 109–182. [Google Scholar] [CrossRef]
  7. Goding, C.R. Motif from neural crest to melanoma: Signal transduction and transcription in the melanocyte lineage. Genes Dev. 2000, 14, 1712–1728. [Google Scholar] [CrossRef] [PubMed]
  8. Sun, X.H.; Copeland, N.G.; Jenkins, N.A.; Baltimore, D. Id proteins Id1 and Id2 selectively inhibit DNA binding by one class of helix-loop-helix proteins. Mol. Cell. Biol. 1991, 11, 5603–5611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Ledent, V.; Vervoort, M. The basic helix-loop-helix protein family: Comparative genomics and phylogenetic analysis. Genome Res. 2001, 11, 754–770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Fisher, A.; Caudy, M. The function of hairy-related bHLH repressor proteins in cell fate decisions. Bioessays 1998, 20, 298–306. [Google Scholar] [CrossRef]
  11. Crozatier, M.; Valle, D.; Dubois, L.; Ibnsouda, S.; Vincent, A. Collier, a novel regulator of Drosophila head development, is expressed in a single mitotic domain. Curr. Biol. 1996, 6, 707–718. [Google Scholar] [CrossRef] [Green Version]
  12. Bailey, P.C.; Martin, C.; Toledo-Ortiz, G.; Quail, P.H.; Huq, E.; Heim, M.A.; Jakoby, M.; Werber, M.; Weisshaar, B. Update on the basic helix-loop-helix transcription factor gene family in Arabidopsis thaliana. Plant Cell 2003, 15, 2497–2501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Heim, M.A.; Jakoby, M.; Werber, M.; Marti, C.; Weisshaar, B.; Bailey, P.C. The basic Helix-Loop-Helix transcription factor family in plants: A genome-wide study of Protein structure and functional diversity. Mol. Biol. Evol. 2003, 20, 735–747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Toledo-Ortiz, G.; Huq, E.; Quail, P.H. The Arabidopsis basic/helix-loop-helix transcription factor family. Plant Cell 2003, 15, 1749–1770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Buck, M.J.; Atchley, W.R. Phylogenetic Analysis of Plant Basic Helix-Loop-Helix Proteins. J. Mol. Evol. 2003, 56, 742–750. [Google Scholar] [CrossRef]
  16. Castelain, M.; Hir, R.L.; Bellini, C. The non-DNA-binding bHLH transcription factor PRE3/bHLH135/ATBS1/TMO7 is involved in the regulation of light signaling pathway in Arabidopsis. Physiologia Plantarum 2012, 145, 450–460. [Google Scholar] [CrossRef]
  17. Abe, H.; Urao, T.; Ito, T.; Seki, M.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Arabidopsis AtMYC2 (bHLH) and AtMYB2 (MYB) function as transcriptional activators in abscisic acid signaling. Plant Cell 2003, 15, 63–78. [Google Scholar] [CrossRef] [Green Version]
  18. Friedrichsen, D.M.; Nemhauser, J.; Muramitsu, T.; Maloof, J.N.; Alonso, J.; Ecker, J.R.; Furuya, M.; Chory, J. Three redundant brassinosteroid early response genes encode putative bHLH transcription factors required for normal growth. Genetics 2002, 162, 1445–1456. [Google Scholar] [CrossRef]
  19. Liu, Y.; Li, X.; Li, K.; Liu, H.; Lin, C. Multiple bHLH Proteins form Heterodimers to Mediate CRY2-Dependent Regulation of Flowering-Time in Arabidopsis. PLoS Genet. 2013, 9, e1003861. [Google Scholar] [CrossRef] [Green Version]
  20. Yao, X.; Cai, Y.; Yu, D.; Liang, G. bHLH104 confers tolerance to cadmium stress in Arabidopsis thaliana. J. Integr. Plant Biol. 2018, 60, 691–702. [Google Scholar] [CrossRef]
  21. Pires, N.; Dolan, L. Origin and diversification of Basic-helix-loop-helix proteins in plants. Mol. Biol. Evol. 2010, 27, 862–874. [Google Scholar] [CrossRef] [Green Version]
  22. Nambara, E.; Marion-Poll, A. Abscisic acid biosynthesis and catabolism. Annu. Rev. Plant Biol. 2005, 56, 165–185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Lefebvre, V.; North, H.; Frey, A.; Sotta, B.; Seo, M.; Okamoto, M.; Nambara, E.; Marion-Poll, A. Functional analysis of Arabidopsis NCED6 and NCED9 genes indicates that ABA synthesized in the endosperm is involved in the induction of seed dormancy. Plant J. 2006, 45, 309–319. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, F.; Zhang, H.; Ding, L.; Soppe, W.J.J.; Xiang, Y. REVERSAL OF RDO5 1, a homolog of rice seed dormancy 4, interacts with bHLH57 and controls ABA biosynthesis and seed dormancy in Arabidopsis. Plant Cell 2020, 32, 1933–1948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Ni, M.; Tepperman, J.M.; Quail, P.H. PIF3, a Phytochrome-interacting factor necessary for normal photoinduced signal transduction, as a novel basic helix-loop-helix protein. Cell 1998, 95, 657–667. [Google Scholar] [CrossRef] [Green Version]
  26. Oh, E.; Zhu, J.; Wang, Z. Interaction between BZR1 and PIF4 integrates brassinosteroid and environmental responses. Nat. Cell Biol. 2012, 14, 802–809. [Google Scholar] [CrossRef] [Green Version]
  27. Oh, E.; Yamaguchi, S.; Kamiya, Y.; Bae, G.; Chung, W.I.; Choi, G. Light activates the degradation of PIL5 protein to promote seed germination through gibberellin in Arabidopsis. Plant J. 2006, 47, 124–139. [Google Scholar] [CrossRef]
  28. Penfield, S.; Josse, E.; Kannangara, R.; Gilday, A.D.; Halliday, K.J.; Graham, I.A. Cold and light control of seed germination through the bHLH transcription factor SPATULA. Curr. Biol. 2005, 15, 1998–2006. [Google Scholar] [CrossRef] [Green Version]
  29. Groszmann, M.; Bylstra, Y.; Lampugnani, E.R.; Smyth, D.R. Regulation of tissue-specific expression of SPATULA, a bHLH gene involved in carpel development, seedling germination, and lateral organ growth in Arabidopsis. J. Exp. Bot. 2010, 61, 1495–1508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Ito, S.; Song, Y.H.; Josephson-Day, A.R.; Miller, R.J.; Breton, G.; Olmstead, R.G.; Imaizumi, T. FLOWERING BHLH transcriptional activators control expression of the photoperiodic flowering regulator CONSTANS in Arabidopsis. Proc. Natl. Acad. Sci. USA 2012, 109, 3582–3587. [Google Scholar] [CrossRef] [Green Version]
  31. Guo, H.; Yang, H.; Mockler, T.; Lin, C. Regulation of flowering time by Arabidopsis photoreceptors. Science 1998, 279, 1360–1363. [Google Scholar] [CrossRef] [Green Version]
  32. Liu, H.; Yu, X.; Li, K.; Klejnot, J.; Yang, H.; Lisiero, D.; Lin, C. Photoexcited CRY2 interacts with CIB1 to regulate transcription and floral initiation in Arabidopsis. Science 2008, 322, 1535–1539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Liu, Y.; Li, Y.; Ma, D.; Chen, Z.; Wang, J.; Liu, H. CIB1 and CO interact to mediate CRY2-dependent regulation of flowering. EMBO Rep. 2018, 19, e45762. [Google Scholar] [CrossRef] [PubMed]
  34. Sharma, N.; Xin, R.; Kim, D.; Sung, S.; Lange, T.; Huq, E. NO FLOWERING IN SHORT DAY (NFL) is a bHLH transcription factor that promotes flowering specifically under short-day conditions in Arabidopsis. Development 2016, 143, 682–690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Wang, H.; Li, Y.; Pan, J.; Lou, D.; Hu, Y.; Yu, D. The bHLH transcription factors MYC2, MYC3, and MYC4 are required for jasmonate-mediated inhibition of flowering in Arabidopsis. Mol. Plant 2017, 10, 1461–1464. [Google Scholar] [CrossRef] [Green Version]
  36. Wang, Z.; Yang, Z.; Li, F. Updates on molecular mechanisms in the development of branched trichome in Arabidopsis and nonbranched in cotton. Plant Biotechnol. J. 2019, 17, 1706–1722. [Google Scholar] [CrossRef] [PubMed]
  37. Kumar, S.V.; Lucyshyn, D.; Jaeger, K.E.; Alós, E.; Alvey, E.; Harberd, N.P.; Wigge, P.A. Transcription factor PIF4 controls the thermosensory activation of flowering. Nature 2012, 484, 242–245. [Google Scholar] [CrossRef] [PubMed]
  38. Heisler, M.G.; Atkinson, A.; Bylstra, Y.H.; Walsh, R.; Smyth, D.R. SPATULA, a gene that controls development of carpel margin tissues in Arabidopsis, encodes a bHLH protein. Development 2001, 128, 1089–1098. [Google Scholar] [CrossRef]
  39. Zhao, H.; Li, X.; Ma, L. Basic helix-loop-helix transcription factors and epidermal cell fate determination in Arabidopsis. Plant Signal. Behav. 2012, 7, 1556–1560. [Google Scholar] [CrossRef] [Green Version]
  40. Menand, B.; Yi, K.; Jouannic, S.; Hoffmann, L.; Ryan, E.; Linstead, P.; Schaefer, D.G.; Dolan, L. An ancient mechanism controls the development of cells with a rooting function on land plants. Science 2007, 316, 1477–1480. [Google Scholar] [CrossRef] [Green Version]
  41. Yi, K.; Menand, B.; Bell, E.; Dolan, L. A basic helix-loop-helix transcription factor controls cell growth and size in root hairs. Nat. Genet. 2010, 42, 264–267. [Google Scholar] [CrossRef]
  42. Bruex, A.; Kainkaryam, R.M.; Wieckowski, Y.; Kang, Y.H.; Bernhardt, C.; Xia, Y.; Zheng, X.; Wang, J.Y.; Lee, M.M.; Benfey, P.; et al. A gene regulatory network for root epidermis cell differentiation in Arabidopsis. PLoS Genet. 2012, 8, e1002446. [Google Scholar] [CrossRef] [Green Version]
  43. Han, X.; Zhang, M.; Yang, M.; Hu, Y. Arabidopsis JAZ proteins interact with and suppress RHD6 transcription factor to regulate jasmonate-stimulated root hair development. Plant Cell 2020, 32, 1049–1062. [Google Scholar] [CrossRef]
  44. Lin, Q.; Ohashi, Y.; Kato, M.; Tsuge, T.; Gu, H.; Qu, L.J.; Aoyama, T. GLABRA2 directly suppresses basic helix-loop-helix transcription factor genes with diverse functions in root hair development. Plant Cell 2015, 27, 2894–2906. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Masucci, J.D.; Rerie, W.G.; Foreman, D.R.; Zhang, M.; Galway, M.E.; Marks, M.D.; Schiefelbein, J.W. The homeobox gene GLABRA2 is required for position-dependent cell differentiation in the root epidermis of Arabidopsis thaliana. Development 1996, 122, 1253–1260. [Google Scholar] [CrossRef] [PubMed]
  46. Bernhardt, C.; Lee, M.M.; Gonzalez, A.; Zhang, F.; Lloyd, A.; Schiefelbein, J. The bHLH genes GLABRA3 (GL3) and ENHANCER OF GLABRA3 (EGL3) specify epidermal cell fate in the Arabidopsis root. Development 2003, 130, 6431–6439. [Google Scholar] [CrossRef] [Green Version]
  47. Ramsay, N.A.; Glover, B.J. MYB-bHLH-WD40 protein complex and the evolution of cellular diversity. Trends Plant Sci. 2005, 10, 63–70. [Google Scholar] [CrossRef]
  48. Qi, T.; Huang, H.; Song, S.; Xie, D. Regulation of jasmonate-mediated stamen development and seed production by a bHLH-MYB complex in Arabidopsis. Plant Cell 2015, 27, 1620–1633. [Google Scholar] [CrossRef] [Green Version]
  49. Cui, Y.; Chen, C.L.; Cui, M.; Zhou, W.J.; Wu, H.L.; Ling, H.Q. Four IVa bHLH transcription factors are novel interactors of FIT and mediate JA inhibition of iron uptake in Arabidopsis. Mol. Plant 2018, 11, 1166–1183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Yuan, Y.; Wu, H.; Wang, N.; Li, J.; Zhao, W.; Du, J.; Wang, D.; Ling, H.Q. FIT interacts with AtbHLH38 and AtbHLH39 in regulating iron uptake gene expression for iron homeostasis in Arabidopsis. Cell Res. 2008, 18, 385–397. [Google Scholar] [CrossRef] [PubMed]
  51. Wang, H.Y.; Klatte, M.; Jakoby, M.; Bäumlein, H.; Weisshaar, B.; Bauer, P. Iron deficiency-mediated stress regulation of four subgroup Ib BHLH genes in Arabidopsis thaliana. Planta 2007, 226, 897–908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Colangelo, E.P.; Guerinot, M.L. The essential basic helix-loop-helix protein FIT1 is required for the iron deficiency response. Plant Cell 2004, 16, 3400–3412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Jakoby, M.; Wang, H.Y.; Reidt, W.; Weisshaar, B.; Bauer, P. FRU (BHLH029) is required for induction of iron mobilization genes in Arabidopsis thaliana. FEBS Lett. 2004, 577, 528–534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Yuan, Y.X.; Zhang, J.; Wang, D.W.; Ling, H.Q. AtbHLH29 of Arabidopsis thaliana is a functional ortholog of tomato FER involved in controlling iron acquisition in strategy I plants. Cell Res. 2005, 15, 613–621. [Google Scholar] [CrossRef] [Green Version]
  55. Pitts, R.J.; Cernac, A.; Estelle, M. Auxin and ethylene promote root hair elongation in Arabidopsis. Plant J. 1999, 16, 553–560. [Google Scholar] [CrossRef] [PubMed]
  56. Wu, H.; Chen, C.; Du, J.; Liu, H.; Cui, Y.; Zhang, Y.; He, Y.; Wang, Y.; Chu, C.; Feng, Z.; et al. Co-overexpression FIT with AtbHLH38 or AtbHLH39 in Arabidopsis-enhanced cadmium tolerance via increased cadmium sequestration in roots and improved iron homeostasis of shoots. Plant Physiol. 2012, 158, 790–800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Zhang, J.; Liu, B.; Li, M.; Jin, H.; Wang, P.; Liu, J.; Xiong, F.; Wang, J.; Wang, H. The bHLH transcription factor bHLH104 interacts with IAA-LEUCINE RESISTANT3 and modulates iron homeostasis in Arabidopsis. Plant Cell 2015, 27, 787–805. [Google Scholar] [CrossRef] [Green Version]
  58. Liang, G.; Zhang, H.; Li, X.; Ai, Q.; Yu, D. bHLH transcription factor bHLH115 regulates iron homeostasis in Arabidopsis thaliana. J. Exp. Bot. 2017, 68, 1743–1755. [Google Scholar] [CrossRef]
  59. Tissot, N.; Robe, K.; Gao, F.; Grant-Grant, S.; Boucherez, J.; Bellegrade, F.; Maghiaoui, A.; Marcelin, R.; Izquierdo, E.; Benhamed, M.; et al. Transcriptional integration of the responses to iron availability in Arabidopsis by the bHLH factor ILR3. N. Phytol. 2019, 223, 1433–1446. [Google Scholar] [CrossRef]
  60. Long, T.A.; Tsukagoshi, H.; Busch, W.; Lahner, B.; Salt, D.E.; Benfey, P.N. The bHLH transcription factor POPEYE regulates response to iron deficiency in Arabidopsis roots. Plant Cell 2010, 22, 2219–2236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Selote, D.; Samira, R.; Matthiadis, A.; Gillikin, J.W.; Long, T.A. Iron-binding E3 ligase mediates iron response in plants by targeting basic helix-loop-helix transcription factors. Plant Physiol. 2015, 167, 273–286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Gao, F.; Dubos, C. Transcriptional integration of plant responses to iron availability. J. Exp. Bot. 2021, 72, 2056–2070. [Google Scholar] [CrossRef] [PubMed]
  63. Martín-Barranco, A.; Spielmann, J.; Dubeaux, G.; Vert, G.; Enric Zelazny, E. Dynamic control of high-affinity iron uptake complex in root epidermal cells. Plant Physiol. 2020, 184, 1236–1250. [Google Scholar] [CrossRef]
  64. Tanabe, N.; Noshi, M.; Mori, D.; Nozawa, K.; Tamoi, M.; Shigeoka, S. The basic helix-loop-helix transcription factor, bHLH11 functions in the iron-uptake system in Arabidopsis thaliana. J. Plant Res. 2018, 132, 93–105. [Google Scholar] [CrossRef] [PubMed]
  65. Kurbidaeva, A.; Ezhova, T.; Novokreshchenova, M. Arabidopsis thaliana ICE2 gene: Phylogeny, structural evolution and functional diversification from ICE1. Plant Sci. 2014, 229, 10–22. [Google Scholar] [CrossRef]
  66. Guan, Q.; Wu, J.; Yue, X.; Zhang, Y.; Zhu, J. A Nuclear calcium-sensing pathway is critical for gene regulation and salt stress tolerance in Arabidopsis. PLoS Genet. 2013, 9, e1003755. [Google Scholar] [CrossRef] [Green Version]
  67. Liu, Q.; Tai, H.; Li, S.; Gao, W.; Zhao, M.; Xie, C.; Li, W. bHLH122 is important for drought and osmotic stress resistance in Arabidopsis and in the repression of ABA catabolism. N. Phytol. 2014, 201, 1192–1204. [Google Scholar] [CrossRef]
  68. Balazadeh, S.; Siddiqui, H.; Allu, A.D.; Matallana-Ramirez, L.P.; Caldana, C.; Mehrnia, M.; Zanor, M.; Köhler, B.; Mueller-Roeber, B. A gene regulatory network controlled by the NAC transcription factor ANAC092/AtNAC2/ORE1 during salt-promoted senescence. Plant J. 2010, 62, 250–264. [Google Scholar] [CrossRef]
  69. Wu, A.; Allu, A.D.; Garapati, P.; Siddiqui, H.; Dortay, H.; Zanor, M.-I.; Asensi-Fabado, M.A.; Munne’-Bosch, S.; Antonio, C.; Tohge, T.; et al. JUNGBRUNNEN1, a reactive oxygen species-responsive NAC transcription factor, regulates longevity in Arabidopsis. Plant Cell 2012, 24, 482–506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Sakuraba, Y.; Bülbül, S.; Piao, W.; Choi, G.; Paek, N. Arabidopsis EARLY FLOWERING3 increases salt tolerance by suppressing salt stress response pathways. Plant J. 2017, 92, 1106–1120. [Google Scholar] [CrossRef] [Green Version]
  71. Kim, J.; Yi, H.; Choi, G.; Shin, B.; Song, P.S.; Choi, G. Functional characterization of phytochrome interacting factor 3 in phytochrome-mediated light signal transduction. Plant Cell 2003, 15, 2399–2407. [Google Scholar] [CrossRef] [Green Version]
  72. Jiang, B.; Shi, Y.; Zhang, X.; Xin, X.; Qi, L.; Guo, H.; Li, J.; Yang, S. PIF3 is a negative regulator of the CBF pathway and freezing tolerance in Arabidopsis. Proc. Natl. Acad. Sci. USA 2017, 114, E6695–E6702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Paik, I.; Kathare, P.K.; Kim, J.; Huq, E. Expanding roles of PIFs in signal integration from multiple processes. Mol. Plant 2017, 10, 1035–1046. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Huq, E.; Quail, P.H. PIF4, a phytochrome-interacting bHLH factor, functions as a negative regulator of phytochrome B signaling in Arabidopsis. EMBO J. 2002, 21, 2441–2450. [Google Scholar] [CrossRef] [Green Version]
  75. Fairchild, C.D.; Schumaker, M.A.; Quail, P.H. HFR1 encodes an atypical bHLH protein that acts in phytochrome A signal transduction. Genes Dev. 2000, 14, 2377–2391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Roig-Villanova, I.; Bou, J.; Sorin, C.; Devlin, P.F.; Martínez-García, J.F. Identification of primary target genes of phytochrome signaling: Early transcriptional control during shade avoidance responses in Arabidopsis. Plant Physiol. 2006, 141, 85–96. [Google Scholar] [CrossRef] [Green Version]
  77. Roig-Villanova, I.; Bou-Torrent, J.; Galstyan, A.; Carretero-paulet, L.; Portolés, S.; Rodríguez-Concepción, M.; Martínez-García, J. Interaction of shade avoidance and auxin response: A role for two novel atypical bHLH proteins. EMBO J. 2007, 26, 4756–4767. [Google Scholar] [CrossRef] [Green Version]
  78. Hao, Y.; Oh, E.; Choi, G.; Liang, Z.; Wang, Z.Y. Interactions between HLH and bHLH factors modulate light-regulated plant development. Mol. Plant 2012, 5, 688–697. [Google Scholar] [CrossRef] [Green Version]
  79. Oh, E.; Kim, J.; Park, E.; Kim, J.; Kang, C.; Choi, G. PIL5, a phytochrome-interacting basic helix-loop-helix protein, is a key negative regulator of seed germination in Arabidopsis thaliana. Plant Cell 2004, 16, 3045–3058. [Google Scholar] [CrossRef] [Green Version]
  80. Oh, E.; Yamaguchi, S.; Hu, J.; Yusuke, J.; Jung, B.; Paik, I.; Lee, H.; Sun, T.; Kamiya, Y.; Choi, G. PIL5, a phytochrome-interacting bHLH protein, regulates gibberellin responsiveness by binding directly to the GAI and RGA promoters in Arabidopsis Seeds. Plant Cell 2007, 19, 1192–1208. [Google Scholar] [CrossRef] [Green Version]
  81. Huq, E.; Al-Sady, B.; Hudson, M.; Kim, C.; Apel, K.; Quail, P.H. PHYTOCHROME-INTERACTING FACTOR 1 is a critical bHLH regulator of chlorophyll biosynthesis. Science 2004, 305, 1937–1941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Moon, J.; Zhu, L.; Shen, H.; Huq, E. PIF1 directly and indirectly regulates chlorophyll biosynthesis to optimize the greening process in Arabidopsis. Proc. Natl. Acad. Sci USA 2008, 105, 9433–9438. [Google Scholar] [CrossRef] [Green Version]
  83. Shen, Y.; Khanna, R.; Carle, C.M.; Quail, P.H. Phytochrome induces rapid PIF5 phosphorylation and degradation in response to red-light activation. Plant Physiol. 2007, 145, 1043–1051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Khanna, R.; Huq, E.; Kikis, E.A.; Al-Sady, B.; Lanzatella, C.; Quail, P.H. A novel molecular recognition motif necessary for targeting photoactivated phytochrome signaling to specific basic helix-loop-helix transcription factors. Plant Cell 2004, 16, 3033–3044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Leivar, P.; Monte, E.; Al-Sady, B.; Carle, C.; Storer, A.; Alonso, J.M.; Ecker, J.R.; Quail, P.H. The Arabidopsis phytochrome-interacting factor PIF7, together with PIF3 and PIF4, regulates responses to prolonged red light by modulating phyB levels. Plant Cell 2008, 20, 337–352. [Google Scholar] [CrossRef] [Green Version]
  86. Fernández-Calvo, P.; Chini, A.; Fernández-Barbero, G.; Chico, J.M.; Gimenez-lbanez, S.; Geerinck, J.; Eeckhout, D.; Schweizer, F.; Godoy, M.; José Manuel Franco-Zorrilla, J.M.; et al. The Arabidopsis bHLH transcription factors MYC3 and MYC4 are targets of JAZ repressors and act additively with MYC2 in the activation of jasmonate responses. Plant Cell 2011, 23, 701–715. [Google Scholar] [CrossRef] [Green Version]
  87. Schweizer, F.; Fernández-Calvo, P.; Zander, M.; Diez-Diaz, M.; Fonseca, S.; Glauser, G.; Lewsey, M.G.; Ecker, J.R.; Solano, R.; Reymond, P. Arabidopsis basic helix-loop-helix transcription factors MYC2, MYC3, and MYC4 regulate glucosinolate biosynthesis, insect performance, and feeding behavior. Plant Cell 2013, 25, 3117–3132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Gonzalez, A.; Zhao, M.; Leavitt, J.M.; Lloyd, A.M. Regulation of the anthocyanin biosynthetic pathway by the TTG1/bHLH/Myb transcriptional complex in Arabidopsis seedlings. Plant J. 2008, 53, 814–827. [Google Scholar] [CrossRef]
  89. Zhao, M.; Morohashi, K.; Hatlestad, G.; Grotewold, E.; Lloyd, A. The TTG1-bHLH-MYB complex controls trichome cell fate and patterning through direct targeting of regulatory loci. Development 2008, 135, 1991–1999. [Google Scholar] [CrossRef] [Green Version]
  90. Qi, T.; Song, S.; Ren, Q.; Wu, D.; Huang, H.; Chen, Y.; Fan, M.; Peng, W.; Ren, C.; Xie, D. The Jasmonate-ZIM-Domain proteins interact with the WD-Repeat/bHLH/MYB complexes to regulate jasmonate-mediated anthocyanin accumulation and trichome initiation in Arabidopsis thaliana. Plant Cell 2011, 23, 1795–1814. [Google Scholar] [CrossRef] [Green Version]
  91. Qi, T.; Huang, H.; Wu, D.; Yan, J.; Qi, Y.; Song, S.; Xie, D. Arabidopsis DELLA and JAZ proteins bind the WD-Repeat/bHLH/MYB complex to modulate gibberellin and jasmonate signaling synergy. Plant Cell 2014, 26, 1118–1133. [Google Scholar] [CrossRef] [Green Version]
  92. Li, X.; Zhang, H.; Ai, Q.; Liang, G.; Yu, D. Two bHLH transcription factors, bHLH34 and bHLH104, regulate iron homeostasis in Arabidopsis thaliana. Plant Physiol. 2016, 170, 2478–2493. [Google Scholar] [CrossRef] [Green Version]
  93. Teale, W.D.; Paponov, I.A.; Palme, K. Auxin in action: Signalling, transport and the control of plant growth and development. Nat. Rev. Mol. Cell Biol. 2006, 7, 847–859. [Google Scholar] [CrossRef] [PubMed]
  94. Varaud, E.; Brioudes, F.; Szécsi, J.; Leroux, J.; Brown, S.; Perrot-Rechenmann, C.; Bendahmane, M. AUXIN RESPONSE FACTOR8 regulates Arabidopsis petal growth by interacting with the bHLH transcription factor BIGPETALp. Plant Cell 2011, 23, 973–983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Zheng, K.; Wang, Y.; Zhang, N.; Jia, Q.; Wang, X.; Hou, C.; Chen, J.; Wang, S. Involvement of PACLOBUTRAZOL RESISTANCE6/KIDARI, an atypical bHLH transcription factor, in auxin responses in Arabidopsis. Front. Plant Sci. 2017, 8, 1813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Liu, Y.; Ji, X.; Nie, X.; Qu, M.; Zheng, L.; Tan, Z.; Zhao, H.; Huo, L.; Liu, S.; Zhang, B.; et al. Arabidopsis AtbHLH112 regulates the expression of genes involved in abiotic stress tolerance by binding to their E-box and GCG-box motifs. N. Phytol. 2015, 207, 692–709. [Google Scholar] [CrossRef] [PubMed]
  97. Wang, W.; Zhu, J.; Lu, Y. Overexpression of AtbHLH112 suppresses lateral root emergence in Arabidopsis. Funct. Plant Biol. 2014, 41, 342–352. [Google Scholar] [CrossRef] [Green Version]
  98. Kim, J.; Kim, H.Y. Molecular characterization of a bHLH transcription factor involved in Arabidopsis abscisic acid-mediated response. BBA Gene Struct. Expr. 2006, 1759, 191–194. [Google Scholar] [CrossRef]
  99. Li, H.; Sun, J.; Xu, Y.; Jiang, H.; Wu, X.; Li, C. The bHLH-type transcription factor AtAIB positively regulates ABA response in Arabidopsis. Plant Mol. Biol. 2007, 65, 655–665. [Google Scholar] [CrossRef] [PubMed]
  100. Tian, H.; Guo, H.; Dai, X.; Cheng, Y.; Zheng, K.; Wang, X.; Wang, S. An ABA down-regulated bHLH transcription repressor gene, bHLH129 regulates root elongation and ABA response when overexpressed in Arabidopsis. Sci. Rep. 2015, 5, 17587. [Google Scholar] [CrossRef] [Green Version]
  101. Zheng, K.; Wang, Y.; Wang, S. The non-DNA binding bHLH transcription factor Paclobutrazol Resistances are involved in the regulation of ABA and salt responses in Arabidopsis. Plant Physiol. Biochem. 2019, 139, 239–245. [Google Scholar] [CrossRef] [PubMed]
  102. Moreno, J.E.; Moreno-Piovano, G.; Chan, R.L. The antagonistic basic helix-loop-helix partners BEE and IBH1 contribute to control plant tolerance to abiotic stress. Plant Sci. 2018, 271, 143–150. [Google Scholar] [CrossRef]
  103. Jiang, Y.; Yang, B.; Deyholos, M.K. Functional characterization of the Arabidopsis bHLH92 transcription factor in abiotic stress. Mol. Genet. Genomics. 2009, 282, 503–516. [Google Scholar] [CrossRef]
  104. Bai, M.Y.; Fan, M.; Oh, E.; Wang, Z.Y. A triple helix-loop-helix/basic helix-loop-helix cascade controls cell elongation downstream of multiple hormonal and environmental signaling pathways in Arabidopsis. Plant Cell 2012, 24, 4917–4929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Oh, E.; Zhu, J.; Bai, M.Y.; Arenhart, R.A.; Sun, Y.; Wang, Z.Y. Cell elongation is regulated through a central circuit of interacting transcription factors in the Arabidopsis hypocotyl. eLife 2014, 3, e03031. [Google Scholar] [CrossRef]
  106. Mara, C.D.; Huang, T.; Irish, V.F. The Arabidopsis floral homeotic proteins APETALA3 and PISTILLATA negatively regulate the BANQUO genes implicated in light signaling. Plant Cell 2010, 22, 690–702. [Google Scholar] [CrossRef] [Green Version]
  107. Zhang, L.; Bai, M.Y.; Wu, J.; Zhu, J.; Wang, H.; Zhang, Z.; Wang, W.; Sun, Y.; Zhao, J.; Sun, X.; et al. Antagonistic HLH/bHLH transcription factors mediate brassinosteroid regulation of cell elongation and plant development in rice and Arabidopsis. Plant Cell 2009, 21, 3767–3780. [Google Scholar] [CrossRef] [Green Version]
  108. Ikeda, M.; Mitsuda, N.; Ohme-Takagi, M. ATBS1 INTERACTING FACTORs negatively regulate Arabidopsis cell elongation in the triantagonistic bHLH system. Plant Signal. Behav. 2013, 8, e23448. [Google Scholar] [CrossRef] [Green Version]
  109. Wang, F.; Gao, Y.; Liu, Y.; Zhang, X.; Gu, X.; Ma, D.; Zhao, Z.; Yuan, Z.; Xue, H.; Liu, H. BES1-regulated BEE1 controls photoperiodic flowering downstream of blue light signaling pathway in Arabidopsis. N. Phytol. 2019, 223, 1407–1419. [Google Scholar] [CrossRef]
  110. Malinovsky, F.G.; Batoux, M.; Schwessinger, B.; Youn, J.H.; Stransfeld, L.; Win, J.; Kim, S.K.; Zipfel, C. Antagonistic regulation of growth and immunity by the arabidopsis basic helix-loop-helix transcription factor HOMOLOG OF BRASSINOSTEROID ENHANCED EXPRESSION2 INTERACTING WITH INCREASED LEAF INCLINATION1 BINDING bHLH1. Plant Physiol. 2014, 164, 1443–1455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Yin, Y.; Vafeados, D.; Tao, Y.; Yoshida, S.; Asami, T.; Chory, J. A new class of transcription factors mediates brassinosteroid-regulated gene expression in Arabidopsis. Cell 2005, 120, 249–259. [Google Scholar] [CrossRef] [Green Version]
  112. Cifuentes-Esquivel, N.; Bou-Torrent, J.; Galstyan, A.; Gallemí, M.; Sessa, G.; Salla Martret, M.; Roig-Villanova, I.; Ruberti, I.; Martínez-García, J.F. The bHLH proteins BEE and BIM positively modulate the shade avoidance syndrome in Arabidopsis seedlings. Plant J. 2013, 75, 989–1002. [Google Scholar] [CrossRef] [PubMed]
  113. Song, Y.; Yang, C.; Gao, S.; Zhang, W.; Li, L.; Kuai, B. Age-triggered and dark-induced leaf senescence require the bHLH transcription factors PIF3, 4, and 5. Mol. Plant 2014, 7, 1776–1787. [Google Scholar] [CrossRef] [Green Version]
  114. Liu, Y.; Lin-Wang, K.; Espley, R.V.; Wang, L.; Yang, H.; Yu, B.; Dare, A.; Varkonyi-Gasic, E.; Wang, J.; Zhang, J.; et al. Functional diversification of the potato R2R3 MYB anthocyanin activators AN1, MYBA1, and MYB113 and their interaction with basic helix-loop-helix cofactors. J. Exp. Bot. 2016, 67, 2159–2176. [Google Scholar] [CrossRef] [PubMed]
  115. Zhao, L.; Gao, L.; Wang, H.; Chen, X.; Wang, Y.; Yang, H.; Wei, C.; Wan, X.; Xia, T. The R2R3-MYB, bHLH, WD40, and related transcription factors in flavonoid biosynthesis. Funct. Integr. Genomic 2013, 13, 75–98. [Google Scholar] [CrossRef]
  116. Lepiniec, L.; Debeaujon, I.; Routaboul, J.; Baudry, A.; Pourcel, L.; Nesi, N.; Caboche, M. Genetics and biochemistry of seed flavonoids. Annu Rev. Plant Biol. 2006, 57, 405–430. [Google Scholar] [CrossRef] [PubMed]
  117. Allan, A.C.; Hellens, R.P.; Laing, W.A. MYB transcription factors that colour our fruit. Trends Plant Sci. 2008, 13, 99–102. [Google Scholar] [CrossRef]
  118. Morohashi, K.; Zhao, M.; Yang, M.; Read, B.; Lloyd, A.; Lamb, R.; Grotewold, E. Participation of the Arabidopsis bHLH factor GL3 in trichome initiation regulatory events. Plant Physiol. 2007, 145, 736–746. [Google Scholar] [CrossRef] [Green Version]
  119. Feng, Y.; Xu, P.; Li, B.; Li, P.; Wen, X.; An, F.; Gong, Y.; Xin, Y.; Zhu, Z.; Wang, Y.; et al. Ethylene promoters root hair growth through coordinated EIN3/EIL1 and RHD6/RSL1 activity in Arabidopsis. Proc. Natl. Acad. Sci. USA 2017, 114, 13834–13839. [Google Scholar] [CrossRef] [Green Version]
  120. Ohashi-Ito, K.; Bergmann, D.C. Arabidopsis FAMA controls the final proliferation/differentiation switch during stomatal development. Plant Cell 2006, 18, 2493–2505. [Google Scholar] [CrossRef] [Green Version]
  121. Wang, N.; Cui, Y.; Liu, Y.; Fan, H.; Du, J.; Huang, Z.; Yuan, Y.; Wu, H.; Ling, H. Requirement and functional redundancy of Ib subgroup bHLH proteins for iron deficiency responses and uptake in Arabidopsis thaliana. Mol. Plant 2013, 6, 503–513. [Google Scholar] [CrossRef] [Green Version]
  122. Lei, R.; Li, Y.; Cai, Y.; Li, C.; Pu, M.; Lu, C.; Yang, Y.; Liang, G. bHLH121 functions as a direct link that facilitates the activation of FIT by bHLH IVc transcription factors for maintaining Fe homestasis in Arabidopsis. Mol. Plant 2020, 13, 634–649. [Google Scholar] [CrossRef]
  123. Ivanov, R.; Brumbarova, T.; Bauer, P. Fitting into the harsh reality: Regulation of iron-deficiency responses in dicotyledonous plants. Mol. Plant 2012, 5, 27–42. [Google Scholar] [CrossRef] [Green Version]
  124. Matthiadis, A.; Long, T.A. Further insight into BRUTUS domain composition and functionality. Plant Signal. Behav. 2016, 11, 8. [Google Scholar] [CrossRef] [Green Version]
  125. Hindt, M.N.; Akmakjian, G.Z.; Pivarski, K.L.; Punshon, T.; Baxter, I.; Salt, D.E.; Guerinot, M.L. BRUTUS and its paralogs, BTS LIKE1 and BTS LIKE2, encode important negative regulators of the iron deficiency response in Arabidopsis thaliana. Metallomics 2017, 9, 876–890. [Google Scholar] [CrossRef] [PubMed]
  126. Eide, D.; Broderius, M.; Fett, J.; Guerinot, M.L. A novel iron-regulated metal transporter from plants identified by functional expression in yeast. Proc. Natl. Acad. Sci. USA. 1996, 93, 5624–5628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Vert, G.; Grotz, N.; Dédaldéchamp, F.; Gaymard, F.; Guerinot, M.L.; Briata, J.F.; Curie, C. IRT1, an Arabidopsis transporter essential for iron uptake from the soil and for plant growth. Plant Cell 2002, 14, 1223–1233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Chinnusamy, V.; Ohta, M.; Kanrar, S.; Lee, B.H.; Hong, X.; Agarwal, M.; Zhu, J.K. ICE1: A regulator of cold-induced transcriptome and freezing tolerance in Arabidopsis. Genes Dev. 2003, 17, 1043–1054. [Google Scholar] [CrossRef] [Green Version]
  129. Jaglo-Ottosen, K.R.; Gilmour, S.J.; Zarka, D.G.; Schabenberger, O.; Thomashow, M.F. Arabidopsis CBF1 overexpression induces COR genes and enhances freezing tolerance. Science 1998, 280, 104–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Liu, Q.; Kasuga, M.; Sakuma, Y.; Abe, H.; Miura, S.; Yamaguchi-Shinozaki, K.; Shinozaki, K. Two transcription factors, DREB1 and DREB2, with an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought- and low-temperature-responsive gene expression, respectively, in Arabidopsis. Plant Cell 1998, 10, 1391–1406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Castillon, A.; Shen, H.; Huq, E. Phytochrome Interacting Factors: Central players in phytochrome-mediated light signaling networks. Trends Plant Sci. 2007, 12, 514–521. [Google Scholar] [CrossRef] [PubMed]
  132. Leivar, P.; Quail, P.H. PIFs: Pivotal components in a cellular signaling hub. Trends Plant Sci. 2011, 16, 19–28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Leivar, P.; Monte, E. PIFs: Systems integrators in plant development. Plant Cell 2014, 26, 56–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Brumbarova, T.; Ivanov, R. The nutrient response transcriptional regulome of Arabidopsis. Iscience 2019, 19, 358–368. [Google Scholar] [CrossRef] [Green Version]
  135. Zhang, X.; Zhu, Z.; An, F.; Hao, D.; Li, P.; Song, J.; Yi, C.; Guo, H. Jasmonate-activated MYC2 represses ETHYLENE INSENSITIVE3 activity to antagonize ethylene-promoted apical hook formation in Arabidopsis. Plant Cell 2014, 26, 1105–1117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Leung, J.; Giraudat, J. Abscisic acid signal transduction. Annu. Rev. Plant Phys. 1998, 49, 199–222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. McCourt, P. Genetic analysis of hormone signaling. Annu Rev. Plant Phys. 1999, 50, 219–243. [Google Scholar] [CrossRef] [Green Version]
  138. Finkelstein, R.R.; Gampala, S.S.; Rock, C.D. Abscisic acid signaling in seeds and seedlings. Plant Cell 2002, 14, S15–S45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Lorenzo, O.; Chico, J.M.; Saénchez-Serrano, J.J.; Solano, R. JASMONATE-INSENSITIVE1 encodes a MYC transcription factor essential to discriminate between different jasmonate-regulated defense responses in Arabidopsis. Plant Cell 2004, 16, 1938–1950. [Google Scholar] [CrossRef] [Green Version]
  140. Anderson, J.P.; Badruzsaufari, E.; Schenk, P.M.; Manners, J.M.; Desmond, O.J.; Ehlert, C.; Maclean, D.J.; Ebert, P.R.; Kazan, K. Antagonistic interaction between abscisic acid and jasmonate-ethylene signaling pathways modulates defense gene expression and disease resistance in Arabidopsis. Plant Cell 2004, 16, 3460–3479. [Google Scholar] [CrossRef] [Green Version]
  141. Yadav, V.; Mallappa, C.; Gangappa, S.N.; Bhatia, S.; Chattopadhyay, S. A basic helix-loop-helix transcription factor in Arabidopsis, MYC2, acts as a repressor of blue light-mediated photomorphogenic growth. Plant Cell 2005, 17, 1953–1966. [Google Scholar] [CrossRef] [Green Version]
  142. Takahashi, Y.; Ebisu, Y.; Kinoshita, T.; Doi, M.; Okuma, E.; Murata, Y.; Shimazaki, K. bHLH transcription factors that facilitate K+ uptake during stomatal opening are repressed by abscisic acid through phosphorylation. Sci. Signal. 2013, 6, ra48. [Google Scholar] [CrossRef] [PubMed]
  143. Takahashi, Y.; Ebisu, Y.; Shimazaki, K. Reconstitution of abscisic acid signaling from the receptor to DNA via bHLH transcription factors. Plant Physiol. 2017, 174, 815–822. [Google Scholar] [CrossRef] [PubMed]
  144. Ruzinova, M.B.; Benezra, R. Id proteins in development, cell cycle and cancer. Trends Cell Biol. 2003, 13, 410–418. [Google Scholar] [CrossRef]
  145. Hyun, Y.; Lee, I. KIDARI, encoding a non-DNA binding bHLH protein, represses light signal transduction in Arabidopsis thaliana. Plant Mol. Biol. 2006, 61, 283–296. [Google Scholar] [CrossRef] [PubMed]
  146. Ikeda, M.; Fujiwara, S.; Mitsuda, N.; Ohme-Takagi, M. A triantagonistic basic helix-loop-helix system regulates cell elongation in Arabidopsis. Plant Cell 2012, 24, 4483–4497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Jaakola, L. New insights into the regulation of anthocyanin biosynthesis in fruit. Trends Plant Sci. 2013, 18, 477–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Xu, W.; Dubos, C.; Lepiniec, L. Transcriptional control of flavonoid biosynthesis by MYB-bHLH-WDR complexes. Trends Plant Sci. 2015, 20, 176–185. [Google Scholar] [CrossRef] [PubMed]
  149. Li, X.; Duan, X.; Jiang, H.; Sun, Y.; Tang, Y.; Yuan, Z.; Guo, J.; Liang, W.; Chen, L.; Yin, J.; et al. Genome-wide analysis of basic/helix-loop-helix transcription factor family in rice and Arabidopsis. Plant Physiol. 2006, 141, 1167–1184. [Google Scholar] [CrossRef] [Green Version]
  150. Jeong, J.; Kim, K.; Kim, M.E.; Kim, H.G.; Heo, G.S.; Park, O.K.; Park, Y.-I.I.; Choi, G.; Oh, E. Phytochrome and ethylene signaling integration in Arabidopsis occurs via the transcriptional regulation of genes co-targeted by PIFs and EIN3. Front. Plant Sci. 2016, 7, 1055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. bHLHs functions in various signaling pathways.
Figure 1. bHLHs functions in various signaling pathways.
Ijms 22 07152 g001
Figure 2. Simplified model for bHLH PREs, MYCs and PIFs, which play central roles in many pathways in Arabidopsis. PIFs bind to their target genes to regulate plant growth and development via manipulation of signaling pathways. PREs and MYCs also participate in core transcription networks together with PIFs to control plant responses to the environment.
Figure 2. Simplified model for bHLH PREs, MYCs and PIFs, which play central roles in many pathways in Arabidopsis. PIFs bind to their target genes to regulate plant growth and development via manipulation of signaling pathways. PREs and MYCs also participate in core transcription networks together with PIFs to control plant responses to the environment.
Ijms 22 07152 g002
Table 1. Functional characterized bHLH protein family.
Table 1. Functional characterized bHLH protein family.
PathwayAGI
Gene Code
Generic NameSynonymFunctions CharacterizedGroupReference
Seed germinationAt4g01460bHLH57 Involved in seed dormancy processIa[22,23]
At2g20180bHLH15PIF1/PIL5Negative regulator of phytochrome-mediated seed germinationVIIa[24,25,26]
At4g36930bHLH24SPTReduced seed dormancyVIIb[27,28]
FloweringAt1g35460bHLH80FBH1Regulate the effect of CO flowering timeIX[29]
At4g34530bHLH63CIB1Promote CRY2-dependent floral initiationXII[30,31,32]
At5g65640bHLH93NFLInvolved in GA mediated control of flowering timeIIIb[33]
At1g32640bHLH6MYC2Required in the JA pathway for regulating flowering timeIIIe[34]
At5g46760bHLH5MYC3Required in the JA pathway for regulating flowering timeIIIe[34]
At4g17880bHLH4MYC4Required in the JA pathway for regulating flowering timeIIIe[34]
At1g18400bHLH44BEE1Regulate photoperiodic floweringXII[35]
At2g43010bHLH9PIF4Accelerate the flowering by activating FT at high temperatureXII[36]
At4g36930bHLH24SPTPlay a role in floral morphogenesis processesVIIa[37]
Cell fateAt1g66470bHLH83RHD6ROOT HAIR DEFECTIVE6VIIb[13,38,39]
At5g37800bHLH86RSL1Partially redundant and involved in root hair developmentVIIIc[40]
At4g33880bHLH85RSL2Partially redundant and involved in root hair developmentVIIIc[40]
At2g14760bHLH84RSL3RHD6-LIKE 3, required for root-hair growthVIIIc[40,41,42]
At1g27740bHLH54RSL4Promote postmitotic cell growth in root-hair cellsVIIIc[40,41,42]
At5g58010bHLH82LRL3Regulate root hair development. XI[40,41,42]
At5g41315bHLH1GL3Trigger the trichome initiation pathwaysIIIf[41,43,44]
At1g63650bHLH2EGL3Regulate trichome and root hair developmentIIIf[45,46]
At5g46830bHLH28MYC5Calcium-binding transcription factor involved in salt stress signalingIIIe[47]
At1g32640bHLH6MYC2Positive regulator of lateral root formation.IIIe[47]
At5g46760bHLH5MYC3Form the bHLH-MYB complex to regulate the stamen developmentIIIe[47]
At4g17880bHLH4MYC4form the bHLH-MYB complex to regulate the stamen developmentIIIe[47]
At3g24140bHLH97FAMAPromote differentiation of stomatal guard cellsIa[48]
Plant mineral nutrition and abiotic stressAt3g56970bHLH38ORG2Regulate the Fe-deficiency responseIb[49]
At3g56980bHLH39ORG3Regulate the Fe-deficiency responseIb[49]
At2g41240bHLH100 A key regulator of iron-deficiency responses Ib[50]
At5g04150bHLH101 A key regulator of iron-deficiency responses Ib[50]
At2g28160bHLH29FITRegulate iron uptake responsesIIIa[49,50,51,52,53]
At3g19860bHLH121URIAct as an essential part of the iron deficiency signaling pathwayIVb[54]
At3g23210bHLH34IDT1Involved in Fe regulation.IVc[55,56,57,58]
At4g14410bHLH104 Positively regulate Fe deficiency responseIVc[55,56,57,58]
At5g54680bHLH105ILR3Plays an important role in Fe homeostasisIVc[55,56,57,58]
At1g51070bHLH115 Involved in response to FeIVc[55,56,57,58]
At3g47640bHLH47PYERegulate response to iron deficiency in Arabidopsis rootsIVb[59,60,61,62]
At4g36060bHLH11 Basic helix-loop-helix (bHLH) DNA-binding superfamily proteinIVb[63]
Stress responseAt3g26744bHLH116ICE1Improve cold tolerance through an ABA independent pathwayIIIb[64]
At3g06590bHLH148AIF2/RITF1Involved in the detoxification of ROS which generated by salt stressOrphans[65]
At1g61660bHLH122 Mediate multiple response to improve stress toleranceIX[66,67]
At2g43010bHLH9PIF4Accelerate the flowering by activating FT at high temperatureXII[68,69,70]
Light signalingAt2g46970bHLH124PIL1Associated with APRR1/TOC1 and is a member of PIF3 familyVIIa[24,25]
At3g59060bHLH65PIF5/PIL6Involved in shade avoidanceVIIa[24,25]
At3g62090bHLH132PIF6/PIL2Associated with APRR1/TOC1 and is a member of PIF3 familyVIIa[24,25]
At1g09530bHLH8PIF3Interact with photoreceptors phyA and phyB. VIIa[24,71,72]
At2g43010bHLH9PIF4Interact with active PhyB proteinVIIa[24,25,73,74]
At1g02340bHLH26HFR1Involved in phytochrome signalingVIIb[75]
At2g42870bHLH165PAR1Control plant development and as a negative regulator of SASOrphans[76,77,78]
At3g58850bHLH166PAR2Control plant development and as a negative regulator of SASOrphans[76,77,78]
At2g20180bHLH15PIF1/PIL5A key negative regulator of phytochrome-mediated responseVIIa[79,80,81,82]
At5g61270bHLH72PIF7Interacts specifically with Pfr form of phyBVIIb[83,84,85]
JA signalingAt1g32640bHLH6MYC2Regulates diverse JA-dependent functionsIIIe[86,87]
At5g46760bHLH5MYC3Act together with MYC2 and MYC4 to activate JA-responsesIIIe[86,87]
At4g17880bHLH4MYC4Act together with MYC2 and MYC3 to activate JA-responsesIIIe[86,87]
At5g41315bHLH1GL3Repressed by JAZsIIIf[88,89,90,91]
At2g22750bHLH18 Induced by JA and inhibit the transcription of the FITIVa[92]
At2g22760bHLH19 Induced by JA and inhibit the transcription of the FITIVa[92]
At2g22770bHLH20NAI1Induced by JA and inhibit the transcription of the FITIVa[92]
At4g37850bHLH25 Induced by JA and inhibit the transcription of the FITIVa[92]
IAA signalingAt1g59640bHLH31BPEpInvolved in the control of petal sizeXII[93,94]
At1g26945bHLH163PRE6Involved in ABA and salts responsesXV[93,95]
ABA signalingAt1g61660bHLH112 Mediate multiple response to improve stress toleranceX[96,97]
At2g46510bHLH17AIBInvolved in response to ABA, repress MYC2-activated leaf senescenceIIId[98,99]
At1g32640bHLH6MYC2Induced by dehydration stress, ABA and blue light IIIe[17]
At2g43140bHLH129 Regulate root elongation and ABA response.IX[100]
PREsInvolved in the regulation of ABA and salt responsesXV[101]
BEEsRepressed by ABA and responses to abiotic stressVa[102]
At5g43650bHLH92 Involved in salt and osmotic stress toleranceIVd[103]
Cross-talk between light and phyto-hormonesAt5g39860bHLH136PRE1/BNQ1Mediate brassinosteroid regulation of cell elongationXV[25,104,105]
At5g15160bHLH134PRE2/BNQ3PHD finger family proteinXV[95]
At3g47710bHLH161PRE4/BNQ3Required for appropriate regulation of flowering time and regulating light responses.XV[106]
At3g28857bHLH164PRE5Involved in the regulation of the light, GA, BR signaling pathwaysXV[25,104]
At1g26945bHLH163PRE6/KIDARIInteracts with HFR1 and negatively regulates its activity.XV[95]
At2g43060bHLH158IBH1ILI1 binding bHLH 1Orphans[107,108]
At2g18300bHLH64HBI1Involved in positive regulation of cell elongation and proliferationXII[104]
At2g43010bHLH9PIF4Negatively regulate phyB mediated responses and involved in SAS VIIa[25,104]
At1g18400bHLH44BEE1Function in the early response to BRsXII[18,109]
At4g36540bHLH58BEE2Function redundant with BEE1/3XII[18,110]
At1g73830bHLH50BEE3Function redundant with BEE1/2XII[18]
At5g08130bHLH46BIM1BES1-INTERACTING MYC-LIKE 1, involved in BRs signalingVa[111,112]
At1g69010bHLH102BIM2Involved in brassinosteroid signaling and modulated SASVa[111,112]
At5g38860bHLH141BIM3Involved in brassinosteroid signaling and modulated SASVa[111,112]
Other aspectsAt4g16430bHLH3JAM3Repress MYC2-activated leaf senescence, negatively regulate JA responseIIId[47]
At1g01260bHLH13JAM2/MYC7ERepress MYC2-activated leaf senescence, negatively regulate JA responseIIId[47]
At4g00870bHLH14 Repress MYC2-activated leaf senescence, negatively regulate JA responsesIIId[47]
At2g46510bHLH17AIBInvolved in response to ABA, repress MYC2-activated leaf senescenceIIId[47]
PIFsPromoted leaf senescenceVIIa[113]
MYCsPositive regulator of Positively regulates flavonoid biosynthesisIIIe[67,114,115,116,117]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hao, Y.; Zong, X.; Ren, P.; Qian, Y.; Fu, A. Basic Helix-Loop-Helix (bHLH) Transcription Factors Regulate a Wide Range of Functions in Arabidopsis. Int. J. Mol. Sci. 2021, 22, 7152. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22137152

AMA Style

Hao Y, Zong X, Ren P, Qian Y, Fu A. Basic Helix-Loop-Helix (bHLH) Transcription Factors Regulate a Wide Range of Functions in Arabidopsis. International Journal of Molecular Sciences. 2021; 22(13):7152. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22137152

Chicago/Turabian Style

Hao, Yaqi, Xiumei Zong, Pan Ren, Yuqi Qian, and Aigen Fu. 2021. "Basic Helix-Loop-Helix (bHLH) Transcription Factors Regulate a Wide Range of Functions in Arabidopsis" International Journal of Molecular Sciences 22, no. 13: 7152. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22137152

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop